Modulating Sterics in Trimethylplatinum(IV) Diimine Complexes To

Jun 27, 2011 - For example, a number of PtII complexes have been shown to promote stoichiometric and/or ..... Nicolaou, K. C.; Bulger, Paul G.; Sarlah...
0 downloads 0 Views 1MB Size
COMMUNICATION pubs.acs.org/Organometallics

Modulating Sterics in Trimethylplatinum(IV) Diimine Complexes To Achieve CC Bond-Forming Reductive Elimination Michael P. Lanci,†,§ Matthew S. Remy,‡ David B. Lao,† Melanie S. Sanford,*,‡ and James M. Mayer*,† † ‡

Department of Chemistry, University of Washington, Box 351700, Seattle, Washington 98195-1700, United States Department of Chemistry, University of Michigan, 930 North University Avenue, Ann Arbor, Michigan 48109, United States

bS Supporting Information ABSTRACT: Tuning the aryl substituents of N,N0 -diaryl-2,3-dimethyl-1,4-diaza1,3-butadiene (DAB) ligands promotes the challenging CC bond-forming reductive elimination from PtIV diimine complexes [(DAB)Pt(CH3)3(solvent)]+ (2) under mild conditions. Experimental results and density functional calculations indicate that 2,6-aryl substitution promotes reductive elimination by facilitating dissociation of the coordinating solvent by close to 10 kcal mol1, but too much steric bulk inhibits the formation of 2 in the one-electron outersphere oxidation of (DAB)Pt(CH3)2 (1).

T

ransition-metal-catalyzed CH activation/CC bondforming reactions provide atom-economical routes for the synthesis of commodity chemicals, pharmaceuticals, natural products, and materials.1 Palladium-based catalysts have been widely used for such transformations, in part because CC bond-forming reductive elimination generally proceeds efficiently from Pd centers.2 The development of related Pt-catalyzed reactions is attractive because PtII is well-known to effect the CH activation of substrates that are challenging using analogous PdII species. For example, a number of PtII complexes have been shown to promote stoichiometric and/or catalytic activation of simple alkanes, including methane.3 However, despite significant efforts in this area, Pt-catalyzed alkane CH functionalization reactions to generate new CC bonds have not yet been achieved. One particularly exciting target for Pt-catalyzed CH activation/CC coupling would be the catalytic oxidative dimerization of CH4.4 This transformation would provide a direct method for upgrading methane to ethane, which is a key precursor to ethylene and higher alkanes. A potential catalytic cycle for Pt-catalyzed methane dimerization could start with methane CH activation3 to form a PtIICH3 complex (Scheme 1, step 1). Oxidatively induced disproportionation could then produce a trimethylplatinum(IV) complex, as has been demonstrated by Bercaw5 and Tilset6 for (DAB)PtII(CH3)2 (1; DAB = N,N0 -diaryl-2,3-dimethyl-1,4-diaza1,3-butadiene) and by us for related Pd species7 (Scheme 1, step 2).8 Importantly, similar (DAB)PtII complexes are known to activate methane, although typically only by methyl group exchange.9 The final step of the catalytic cycle would then involve CC bondforming reductive elimination from PtIV to generate ethane (Scheme 1, step 3). Such reductive elimination is typically challenging for PtIV methyl complexes with diimine ligands,10,11 although it is known for related complexes with phosphine or β-diketiminate ligands and for PdIV polymethyl adducts.10,12,13 r 2011 American Chemical Society

Scheme 1. Proposed Catalytic Cycle for Pt-Catalyzed Methane Dimerization

We hypothesized that increased steric bulk in the diimine ligand environment would facilitate reductive elimination from [(DAB)PtIV(CH3)3(solvent)]+ (2). This hypothesis is predicated on the observation that sterically encumbered diimine ligands accelerate CH3I reductive elimination from PtIV;14 furthermore, bulky ligands have been used to promote reductive elimination in other systems.15 In this communication we demonstrate that modification of the DAB ligands and reaction solvent can be used to achieve facile CC reductive elimination of ethane. In addition, we show that the diimine ligand structure also strongly influences the oxidative disproportionation reaction that converts 1 to 2. These studies have facilitated the identification of a DAB ligand substitution pattern that promotes both steps 2 and 3 in Scheme 1 efficiently at room temperature. A series of (DAB)PtII(CH3)2 complexes (1AD) have been prepared with differently substituted aryl rings: 4-methyl (A), Received: June 13, 2011 Published: June 27, 2011 3704

dx.doi.org/10.1021/om200508k | Organometallics 2011, 30, 3704–3707

Organometallics Chart 1. Yields of Reductive-Elimination Products, C2H6 and 3, from Thermolysis of 2 in Acetone-d6 for 8.5 h at 60 °C (Scheme 2)a

a

COMMUNICATION

Scheme 2. Mechanism for the Fluxionality of 2C,D and Their Reductive Elimination of Ethane

The 1,4-diaza-1,3-butadiene ligands AD are shown.

3,5-dimethyl (B), 2,6-dimethyl (C), and 2,6-diisopropyl (D) (shown in Chart 1).16,17 The cationic PtIV solvento complexes [(DAB)PtIV(CH3)3(solv)]+ (2AD) were then generated by treating 1AD with 1 equiv of CH3I followed by AgPF6 in acetone-d6 (eq 1).18,19 The 1H NMR spectra of 2A,B in acetone-d6 at ambient temperature exhibit two distinct PtCH3 resonances in the expected 2:1 ratio. In contrast, 2C,D show a single PtCH3 signal, indicating that the three methyl ligands are in fast exchange.10b This exchange is expected to proceed via the five-coordinate cationic intermediate I-1 (Scheme 2, top).20 As such, these data suggest that I-1 is significantly more accessible in the more sterically crowded complexes 2C,D.

Consistent with literature precedent, compounds 2A,B are quite thermally stable.10,11 For example, no decay was observed after heating for 24 h in acetone-d6 at 60 °C. Even after 18 h at 100 °C, C2H6 was not detected and only minimal (30 kcal mol1. In contrast, the ortho-aryl-substituted diimine complexes 2C,D quantitatively eliminate ethane over 8.5 h at 60 °C in acetone-d6, in 96(6) and 91(6)% yields, respectively (Chart 1).21 This reaction proceeds with concomitant formation of the corresponding PtII monomethyl solvento complexes [(DAB)PtII CH3(solv)]+ in 82(6)% (3C) and 72(5)% (3D) yield (Pt yields are less than quantitative due to some competing decomposition). The kinetics of CC coupling were monitored by 1H NMR spectroscopy. First-order rate constants were observed for the decay of 2C ([1.6(3)]  104 s1) and 2D ([4.2(6)]  104 s1) at 60 °C. These correspond to energy barriers (ΔGq60 °C) of 25.4(1) and 24.7(1) kcal mol1 for 2C,D, respectively.22 DFT calculations were conducted to gain insights into the marked reactivity difference between the complexes with and without 2,6-substituents on the aryl rings. Gas-phase reaction coordinates for reductive elimination from isomeric 2B and 2C

Figure 1. Reaction coordinate for C2H6 elimination (gas-phase free energies in kcal mol1) from DFT calculations.17

(with solv = acetone) were computed using the B3LYP functional in Gaussian 03 and the basis sets LANL2TZ on Pt and 6-31G** on C, H, N, and O (Figure 1).23 Since it is well-known that reductive elimination reactions from six-coordinate PtIV trimethyl complexes proceed via initial ligand dissociation,10,24 we first calculated the free energies for loss of acetone to give I1B or I-1C. This step is 5.5 kcal mol1 uphill for 2B but is roughly isoergic for 2C. These results are consistent with the NMR data discussed above, indicating much more facile loss of solvent from 2C. In the transition structures (TS) for H3CCH3 coupling, the two reacting methyl groups straddle the N2Pt equatorial plane with a C 3 3 3 C distance of ∼1.95 Å. Distortion along the imaginary mode in one direction and reoptimization of the minimum energy geometries returned the five-coordinate square-pyramidal complexes I-1B and I-1C. In the other direction, a slight decrease in the C 3 3 3 C distance led to C2H6-bound complexes I-2B and I-2C as local minima (Scheme 2). Displacement of coordinated C2H6 with acetone then completes the reaction sequence to form 3B or 3C. The overall gas-phase free energy barrier for C2H6 formation from 2C is 21.8 kcal mol1 at 298 K, which is in reasonable agreement with the experimental value (ΔGq60 °C = 25.4 kcal mol1 in acetone solution). The barrier for 2B is computed to be 5.4 kcal mol1 higher than that of 2C, again consistent with experiment (ΔΔGq60 °C > 5.6 kcal mol1). The calculations show that the difference in energy between the transition structures is almost entirely due to ΔΔG° for acetone dissociation (5.5 kcal mol1). Thus, the ortho substituents accelerate reductive elimination by facilitating access to the five-coordinate intermediate I-1, rather than by promoting the actual CC bond-forming step. 3705

dx.doi.org/10.1021/om200508k |Organometallics 2011, 30, 3704–3707

Organometallics Scheme 3. Oxidation of (DAB)PtIIMe2 (2) with AcFc+

The DFT calculations indicate that CC coupling from 2 should be faster in more weakly coordinating solvents.25 Indeed, in nitromethane-d3, reductive elimination to C2H6 from 2C proceeds within 8 h at room temperature (k = [5.7(5)]  105 s1 and ΔGq25 °C = 23.2(1) kcal mol1). The use of CD2Cl2 resulted in quantitative reductive elimination to C2H6 within 1 h at room temperature (k = 7.6(14)  104 s1; ΔGq25 °C = 21.7(1) kcal mol1).26 This order of facility of reductive elimination, dichloromethane > nitromethane > acetone, parallels the lower donor ability of these solvents.27 This is almost as fast as reductive elimination from the 1,2-bis(diphenylphosphino)ethane (dppe) complex (dppe)Pt(CH3)3(OTf), which occurs upon warming to room temperature in acetone (though complexes of this type have not been shown to activate CH bonds).10b Furthermore, although 2A,B are both stable to heating in nitromethane-d3 at 60 °C (